首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Prostaglandin-endoperoxide H synthases (PGHSs) have a cyclooxygenase that forms prostaglandin (PG) G2 from arachidonic acid (AA) plus oxygen and a peroxidase that reduces the PGG2 to PGH2. The peroxidase activates the cyclooxygenase. This involves an initial oxidation of the peroxidase heme group by hydroperoxide, followed by oxidation of Tyr385 to a tyrosyl radical within the cyclooxygenase site. His386 of PGHS-1 is not formally part of either active site, but lies in an extended helix between Tyr385, which protrudes into the cyclooxygenase site, and His388, the proximal ligand of the peroxidase heme. When His386 was substituted with alanine in PGHS-1, the mutant retained <2.5% of the native peroxidase activity, but >20% of the native cyclooxygenase activity. However, peroxidase activity could be restored (10-30%) by treating H386A PGHS-1 with cyclooxygenase inhibitors or AA, but not with linoleic acid; in contrast, mere occupancy of the cyclooxygenase site of native PGHS-1 had no effect on peroxidase activity. Heme titrations indicated that H386A PGHS-1 binds heme less tightly than does native PGHS-1. The low peroxidase activity and decreased affinity for heme of H386A PGHS-1 imply that His386 helps optimize heme binding. Molecular dynamic simulations suggest that this is accomplished in part by a hydrogen bond between the heme D-ring propionate and the N-delta of Asn382 of the extended helix. The structure of the extended helix is, in turn, strongly supported by stable hydrogen bonding between the N-delta of His386 and the backbone carbonyl oxygens of Asn382 and Gln383. We speculate that the binding of cyclooxygenase inhibitors or AA to the cyclooxygenase site of ovine H386A PGHS-1 reopens the constriction in the cyclooxygenase site between the extended helix and a helix containing Gly526 and Ser530 and restores native-like structure to the extended helix. Being less bulky than AA, linoleic acid is apparently unable to reopen this constriction.  相似文献   

2.
Self-inactivation imposes an upper limit on bioactive prostanoid synthesis by prostaglandin H synthase (PGHS). Inactivation of PGHS peroxidase activity has been found to begin with Intermediate II, which contains a tyrosyl radical. The structure of this radical is altered by cyclooxygenase inhibitors, such as indomethacin and flurbiprofen, and by replacement of heme by manganese protoporphyrin IX (forming MnPGHS-1). Peroxidase self-inactivation in inhibitor-treated PGHS-1 and MnPGHS-1 was characterized by stopped-flow spectroscopic techniques and by chromatographic and mass spectrometric analysis of the metalloporphyrin. The rate of peroxidase inactivation was about 0.3 s(-)1 in inhibitor-treated PGHS-1 and much slower in MnPGHS-1 (0.05 s(-)1); as with PGHS-1 itself, the peroxidase inactivation rates were independent of peroxide concentration and structure, consistent with an inactivation process beginning with Intermediate II. The changes in metalloporphyrin absorbance spectra during inactivation of inhibitor-treated PGHS-1 were similar to those observed with PGHS-1 but were rather distinct in MnPGHS-1; the kinetics of the spectral transition from Intermediate II to the next species were comparable to the inactivation kinetics in each case. In contrast to the situation with PGHS-1 itself, significant amounts of heme degradation occurred during inactivation of inhibitor-treated PGHS-1, producing iron chlorin and heme-protein adduct species. Structural perturbations at the peroxidase site (MnPGHS-1) or at the cyclooxygenase site (inhibitor-treated PGHS-1) thus can influence markedly the kinetics and the chemistry of PGHS-1 peroxidase inactivation.  相似文献   

3.
The cyclooxygenase activity of the bifunctional enzyme prostaglandin H(2) synthase-2 (PGHS-2) is the target of non-steroidal anti-inflammatory drugs. Inhibition of the peroxidase activity of PGHS has been less studied. Using Soret absorption changes, the binding of aromatic hydroxamic acids to the peroxidase site of PGHS-2 was examined to investigate the structural determinants of inhibition. Typical of mammalian peroxidases, the K(d) for benzhydroxamic acid (42mM) is much greater than that for salicylhydroxamic acid (475microM). Binding of the hydroxamic acid tepoxalin (25microM) resulted in only minor Soret changes. However, tepoxalin is an efficient reducing cosubstrate, indicating that it is an alternative electron donor rather than an inhibitor of the peroxidase activity. Aromatic hydrazides are metabolically activated inhibitors of peroxidases. 2-Naphthoichydrazide (2-NZH) caused the time- and concentration-dependent inhibition of both PGHS-2 peroxidase and cyclooxygenase activities. H(2)O(2) was required for the inactivation of both PGHS-2 activities and indomethacin (which binds at the cyclooxygenase site) did not affect the peroxidase inhibitory potency of 2-NZH. A series of aromatic hydrazides were found to be potent inhibitors of PGHS-2 peroxidase activity with IC(50) values in the 6-100microM range for 13 of the 18 hydrazides examined. Selective inhibition of PGHS-2 over myeloperoxidase and horseradish peroxidase isozyme C was increased by certain ring substitutions. In particular, a chloro group para to the hydrazide moiety increased the PGHS-2 selectivity relative to both myeloperoxidase and horseradish peroxidase isozyme C.  相似文献   

4.
In neutral solution, 5,6-dihydrocytidine undergoes spontaneous deamination (k25 approximately 3.2 x 10(-5) s(-1)) much more rapidly than does cytidine (k25 approximately 3.0 x 10(-10) s(-1)), with a more favorable enthalpy of activation (DeltaDeltaH# = -8.7 kcal/mol) compensated by a less favorable entropy of activation (TDeltaDeltaS# = -1.8 kcal/mol at 25 degrees C). E. coli cytidine deaminase enhances the rate of deamination of 5,6-dihydrocytidine (kcat/k(non) = 4.4 x 10(5)) by enhancing the entropy of activation (DeltaDeltaH# = 0 kcal/mol; TDeltaDeltaS# = +7.6 kcal/mol, at 25 degrees C). Binding of the competitive inhibitor 3,4,5,6-tetrahydrouridine (THU), a stable analogue of 5,6-dihydrocytidine in the transition state for its deamination, is accompanied by a release of enthalpy (DeltaH = -7.1 kcal/mol, TDeltaDeltaS = +2.2 kcal/mol) that approaches the estimated enthalpy of binding of the actual substrate in the transition state for deamination of 5,6-dihydrocytidine (DeltaH = -8.1 kcal/mol, TDeltaDeltaS = +6.0 kcal/mol). Thus, the shortcomings of THU in capturing all of the binding affinity expected of an ideal transition-state analogue reflect a less favorable entropy of association. That difference may arise from the analogue's inability to displace a water molecule from the "leaving group site" at which ammonia is generated in the normal reaction. The effect on binding of removing the 4-OH group from the transition-state analogue THU, to form 3,4,5,6-tetrahydrozebularine (THZ) (DeltaDeltaH = -2.1 kcal/mol, TDeltaDeltaS = -4.4 kcal/mol), is mainly entropic, consistent with the inability of THZ to displace water from the "attacking group site". These results are consistent with earlier indications [Snider, M. J., and Wolfenden, R. (2001) Biochemistry 40, 11364] that site-bound water plays a prominent role in substrate activation and inhibitor binding by cytidine deaminase.  相似文献   

5.
Enzyme-substrate contacts in the hydrolysis of ester substrates by the cysteine protease papain were investigated by systematically altering backbone hydrogen-bonding and side-chain hydrophobic contacts in the substrate and determining each substrate's kinetic constants. The observed specificity energies [defined as delta delta G obs = -RT ln [(kcat/KM)first/(kcat/KM)second)]] of the substrate backbone hydrogen bonds were -2.7 kcal/mol for the P2 NH and -2.6 kcal/mol for the P1 NH when compared against substrates containing esters at those sites. The observed binding energies were -4.0 kcal/mol for the P2 Phe side chain, -1.0 kcal/mol for the P1' C=O, and -2.3 kcal/mol for the P2' NH. The latter three values probably all significantly underestimate the incremental binding energies. The P2 NH, P2 Phe side-chain, and P1 NH contacts display a strong interdependence, or cooperativity, of interaction energies that is characteristic of enzyme-substrate interactions. This interdependence arises largely from the entropic cost of forming the enzyme-substrate transition state. As favorable contacts are added successively to a substrate, the entropic penalty associated with each decreases and the free energy expressed approaches the incremental interaction energy. This is the first report of a graded cooperative effect. Elucidation of favorable enzyme-substrate contacts remote from the catalytic site will assist in the design of highly specific cysteine protease inhibitors.  相似文献   

6.
Luzhkov VB  Almlöf M  Nervall M  Aqvist J 《Biochemistry》2006,45(36):10807-10814
We report results from microscopic molecular dynamics and free energy perturbation simulations of substrate binding and selectivity for the Escherichia coli high-affinity ammonium transporter AmtB. The simulation system consists of the protein embedded in a model membrane/water surrounding. The calculated absolute binding free energies for the external NH(4)(+) ions are between -5.8 and -7.3 kcal/mol and are in close agreement with experimental data. The apparent pK(a) of the bound NH(4)(+) increases by more than 4 units, indicating a preference for binding ammonium ion and not neutral ammonia. The external binding site is also selective for NH(4)(+) toward monovalent metal cations by 2.4-4.4 kcal/mol. The externally bound NH(4)(+) shows strong electrostatic interactions with the proximal buried Asp160, stabilized in the anionic form, whereas the interactions with the aromatic rings of Phe107 and Trp148, lining the binding cavity, are less pronounced. Simulated mutation of the highly conserved Asp160 to Asn reduces the pK(a) of the bound ammonium ion by approximately 7 units and causes loss of its binding. The calculations further predict that the substrate affinity of E. coli AmtB depends on the ionization state of external histidines. The computed free energies of hypothetical intermediate states related to transfer of NH(3), NH(4)(+), or H(2)O from the external binding site to the first position inside the internal channel pore favor permeation of the neutral species through the channel interior. However, the predicted change in the apparent pK(a) of NH(4)(+) upon translocation from the external site, Am1, to the first internal site, Am2, indicates that ammonium ion becomes deprotonated only when it enters the channel interior.  相似文献   

7.
Snider MJ  Wolfenden R 《Biochemistry》2001,40(38):11364-11371
Kinetic measurements have shown that substantial enthalpy changes accompany substrate binding by cytidine deaminase, increasing markedly as the reaction proceeds from the ground state (1/K(m), DeltaH = -13 kcal/mol) to the transition state (1/K(tx), DeltaH = -20 kcal/mol) [Snider, M. J., et al. (2000) Biochemistry 39, 9746-9753]. In the present work, we determined the thermodynamic changes associated with the equilibrium binding of inhibitors by cytidine deaminase by isothermal titration calorimetry and van't Hoff analysis of the temperature dependence of their inhibition constants. The results indicate that the binding of the transition state analogue 3,4-dihydrouridine DeltaH = -21 kcal/mol), like that of the transition state itself (DeltaH = -20 kcal/mol), is associated with a large favorable change in enthalpy. The significantly smaller enthalpy change that accompanies the binding of 3,4-dihydrozebularine (DeltaH = -10 kcal/mol), an analogue of 3,4-dihydrouridine in which a hydrogen atom replaces this inhibitor's 4-OH group, is consistent with the view that polar interactions with the substrate at the site of its chemical transformation play a critical role in reducing the enthalpy of activation for substrate hydrolysis. The entropic shortcomings of 3,4-dihydrouridine, in capturing all of the free energy involved in binding the actual transition state, may arise from its inability to displace a water molecule that occupies the binding site normally occupied by product ammonia.  相似文献   

8.
Wu G  Kulmacz RJ  Tsai AL 《Biochemistry》2003,42(46):13772-13777
The peroxidase and cyclooxygenase activities of prostaglandin H synthase-1 (PGHS-1) both become irreversibly inactivated during reaction with peroxide. Sequential stopped-flow absorbance measurements with a chromogenic peroxidase cosubstrate previously were used to evaluate the kinetics of peroxidase inactivation during reaction of PGHS-1 with peroxide [Wu, G., et al. (1999) J. Biol. Chem. 274, 9231-7]. This approach has now been adapted to use a chromogenic cyclooxygenase substrate to analyze the detailed kinetics of cyclooxygenase inactivation during reaction of PGHS-1 with several hydroperoxides. In the absence of added reducing cosubstrates, which maximizes the levels of oxidized enzyme intermediates expected to lead to inactivation, cyclooxygenase activity was lost as fast as, or somewhat faster than, peroxidase activity. Cyclooxygenase inactivation kinetics appeared to be sensitive to the structure of the peroxide used. The addition of reducing cosubstrate during reaction of PGHS-1 with peroxide protected the peroxidase activity to a much greater degree than the cyclooxygenase activity. The results suggest a new concept of PGHS inactivation: that distinct damage can occur at the two active sites during side reactions of Intermediate II, which forms during reaction of PGHS with peroxide and which contains two oxidants, a ferryl heme in the peroxidase site, and a tyrosyl free radical in the cyclooxygenase site.  相似文献   

9.
Synthetic polymers of a specific shape and size play major role in drug delivery systems. Dendrimers are unique synthetic macromolecules of nanometer dimensions with a highly branched structure and globular shape with potential applications in gene and drug delivery. We examine the interaction of several dendrimers of different compositions mPEG-PAMAM (G3), mPEG-PAMAM (G4) and PAMAM (G4) with hydrophilic and hydrophobic drugs cisplatin, resveratrol, genistein and curcumin at physiological conditions. FTIR and UV-visible spectroscopic methods as well as molecular modeling were used to analyse drug binding mode, the binding constant and the effects of drug complexation on dendrimer stability and conformation. Structural analysis showed that cisplatin binds dendrimers in hydrophilic mode via Pt cation and polymer terminal NH(2) groups, while curcumin, genistein and resveratrol are located mainly in the cavities binding through both hydrophobic and hydrophilic contacts. The overall binding constants of durg-dendrimers are ranging from 10(2) M(-1) to 10(3) M(-1). The affinity of dendrimer binding was PAMAM-G4>mPEG-PAMAM-G4>mPEG-PAMAM-G3, while the order of drug-polymer stability was curcumin>cisplatin>genistein>resveratrol. Molecular modeling showed larger stability for genisten-PAMAM-G4 (ΔG = -4.75 kcal/mol) than curcumin-PAMAM-G4 ((ΔG = -4.53 kcal/mol) and resveratrol-PAMAM-G4 ((ΔG = -4.39 kcal/mol). Dendrimers might act as carriers to transport hydrophobic and hydrophilic drugs.  相似文献   

10.
Cyclooxygenase-2 (COX-2) enzyme binds to arachidonic acid and releases metabolites that are used to induce pain and inflammation. COX-2 selective inhibitors such as celecoxib, rofecoxib and valdecoxib are currently used to reduce inflammatory response. However, they lack anti-thrombotic activity and hence lead to cardiovascular and renal liabilities apart from gastrointestinal irritation. Therefore, there is still a need to develop more potent COX-2 inhibitors. In this paper, we report the screening of various compounds from the ZINC database (contains 4.6 million small molecule compounds) using the eHiTS (electronic High Throughput Screening) software tool against the COX-2 protein. The strategy employed can be conveniently split into two categories, viz. screening and docking, respectively. Screening was performed using molecular constraints tool to filter compounds with physico-chemical properties similar to the 6COX bound ligand SC-558. The analysis resulted in 1042 Lipinski compliant hits which are docked and scored to identify structurally novel ligands that make similar interactions to those of known ligands or may have different interactions with other parts of the binding site. Our screening approach identified two molecules ZINC00663976 (eHITS score of -7.135 kcal/mol) and ZINC02062094 (eHITS score of -7.242 kcal/mol) from the ZINC database. Their energy scores are better than the 6COX bound co-crystallized ligand SC-558 with an eHiTS score of -6.559 kcal/mol. Both the ligands were docked within the binding pocket forming interactions with Leu352, Phe518, Met522, Val523, Ala527 and Ser353. Visual inspection suggested similar orientation and binding mode for ZINC02062094 with SC-558 ligand. The NH group of the ligand formed hydrogen bond interactions with the backbone NH of Ala527.  相似文献   

11.
Nitric oxide (( small middle dot)NO) plays a central role in vascular homeostasis via regulation of smooth muscle relaxation and platelet aggregation. Although mechanisms for ( small middle dot)NO formation are well known, removal pathways are less well characterized, particularly in cells that respond to ( small middle dot)NO through activation of soluble guanylate cyclase. Herein, we report that ( small middle dot)NO is catalytically consumed by prostaglandin H synthase-1 (PGHS-1) through acting as a reducing peroxidase substrate. With purified ovine PGHS-1, ( small middle dot)NO consumption requires peroxide (LOOH or H(2)O(2)), with a K(m)( (app)) for 15(S)hydroperoxyeicosatetraenoic acid (HPETE) of 3. 27 +/- 0.35 microm. During this, 2 mol ( small middle dot)NO are consumed per mol HPETE, and loss of HPETE hydroperoxy group occurs with retention of the conjugated diene spectrum. Hydroperoxide-stimulated ( small middle dot)NO consumption requires heme incorporation, is not inhibited by indomethacin, and is further stimulated by the reducing peroxidase substrate, phenol. PGHS-1-dependent ( small middle dot)NO consumption also occurs during arachidonate, thrombin, or activation of platelets (1-2 microm.min(-1) for typical plasma platelet concentrations) and prevents ( small middle dot)NO stimulation of platelet soluble guanylate cyclase. Platelet sensitivity to ( small middle dot)NO as an inhibitor of aggregation is greater using a platelet-activating stimulus () that does not cause ( small middle dot)NO consumption, indicating that this mechanism overcomes the anti-aggregatory effects of ( small middle dot)NO. Catalytic consumption of ( small middle dot)NO during eicosanoid synthesis thus represents both a novel proaggregatory function for PGHS-1 and a regulated mechanism for vascular ( small middle dot)NO removal.  相似文献   

12.
Phenylamidine cationic groups linked by a furan ring (furamidine) and related symmetric diamidine compounds bind as monomers in the minor groove of AT sequences of DNA. DB293, an unsymmetric derivative with one of the phenyl rings of furamidine replaced with a benzimidazole, can bind to AT sequences as a monomer but binds more strongly to GC-containing minor-groove DNA sites as a stacked dimer. The dimer-binding mode has high affinity, is highly cooperative and sequence selective. In order to develop a better understanding of the correlation between structural and thermodynamic aspects of DNA molecular recognition, DB293 was used as a model to compare the binding of minor-groove agents with AT and mixed sequence DNA sites. Isothermal titration calorimetry and surface plasmon resonance results clearly show that the binding of DB293 and other related compounds into the minor groove of AT sequences is largely entropy-driven while the binding of DB293 as a dimer into the minor groove of GC-containing sequences is largely enthalpy-driven. At 25 degrees C, for example, the AT binding has DeltaG degrees, DeltaH degrees and TDeltaS degrees values of -9.6, -3.6 and 6.0 kcal/mol while the values for dimer binding to a GC-containing site are -9.0, -10.9 and -1.9 kcal/mol (per mol of bound compound), respectively. These results show that the thermodynamic components for binding of compounds of this type to DNA are very dependent on the structure, solvation and sequence of the DNA binding site.  相似文献   

13.
Epidemiological studies suggest that Mediterranean diets rich in resveratrol are associated with reduced risk of coronary artery disease. However, the mechanisms by which resveratrol exerts its vasculoprotective effects are not completely understood. Because oxidative stress and endothelial cell injury play a critical role in vascular aging and atherogenesis, we evaluated whether resveratrol inhibits oxidative stress-induced endothelial apoptosis. We found that oxidized LDL and TNF-alpha elicited significant increases in caspase-3/7 activity in endothelial cells and cultured rat aortas, which were prevented by resveratrol pretreatment (10(-6)-10(-4) mol/l). The protective effect of resveratrol was attenuated by inhibition of glutathione peroxidase and heme oxygenase-1, suggesting a role for antioxidant systems in the antiapoptotic action of resveratrol. Indeed, resveratrol treatment protected cultured aortic segments and/or endothelial cells against increases in intracellular H(2)O(2) levels and H(2)O(2)-mediated apoptotic cell death induced by oxidative stressors (exogenous H(2)O(2), paraquat, and UV light). Resveratrol treatment also attenuated UV-induced DNA damage (comet assay). Resveratrol treatment upregulated the expression of glutathione peroxidase, catalase, and heme oxygenase-1 in cultured arteries, whereas it had no significant effect on the expression of SOD isoforms. Resveratrol also effectively scavenged H(2)O(2) in vitro. Thus resveratrol seems to increase vascular oxidative stress resistance by scavenging H(2)O(2) and preventing oxidative stress-induced endothelial cell death. We propose that the antioxidant and antiapoptotic effects of resveratrol, together with its previously described anti-inflammatory actions, are responsible, at least in part, for its cardioprotective effects.  相似文献   

14.
Since the discovery of the inducible form of prostaglandin (PG) H synthase (PGHS), PGHS-2, considerable effort has been made to design selective inhibitors of this isozyme. N-(2-cyclohexyloxy-4-nitrophenyl) methanesulfonamide (NS-398) and 5-bromo-2-(4-fluorophenyl)-3-(4-methylsulfonyl) thiophene (DuP-697) have been shown to interact reversibly with PGHS-1, while irreversibly inhibiting PGHS-2 in a time-dependent manner. In the present study we have tested the effects of DuP-697 and NS-398 on the activity of PGHS-1 and further explored the interactions between these agents and the inhibition of PGHS-1 by aspirin, indomethacin and ibuprofen. Three independent experimental systems, namely bovine aortic endothelial cells (BAEC), human fibroblasts and ram seminal vesicle microsomes were used to investigate the effects of DuP-697 and NS-398 on PGHS-1. The results show that DuP-697 and NS-398, at concentrations ranges which do not inhibit PGHS-1 activity, significantly attenuated the inhibition of PGHS-1 that was caused by aspirin and indomethacin. The same concentrations of DuP-697 and NS-398 did not affect the inhibition of PGHS-1 that was induced by the competitive reversible inhibitors ibuprofen and naproxen. Similar effects of DuP-697 and NS-393 were obtained with ram seminal vesicle microsomes. These results suggest that PGHS-2 inhibitors DuP-697 and NS-398 possibly interact with PGHS-1 at a site different from the enzyme's catalytic site, thus causing attenuation of PGHS-1 inhibition by aspirin and indomethacin without altering PGHS-1 basal activity or the ibuprofen-induced inhibition.  相似文献   

15.
Two new parameters, I: and C:, are introduced for the quantitative evaluation of functional chimeras: I: (impact) and C: (context dependence) are the free energy difference and sum, respectively, of the effects on a given property measured in forward and retro chimeras. The forward chimera is made by substitution of a part "a" from ensemble A into the analogous position of homologous ensemble B (S:(B --> A)). The C: value is a measure of the interaction of the interrogated position with its surroundings, whereas I: is an expression of the quantitative importance of the probed position. Both I: and C: vary with the evaluated property, for example, kinetics, binding, thermostability, and so forth. The retro chimera is the reverse substitution of the analogous part "b" from B into A, S:(A --> B). The I: and C: values derived from original data for forward and retro mutations in aspartate and tyrosine aminotransferase, from literature data for quasi domain exchange in oncomodulin and for the interaction of Tat with bovine and human TAR are evaluated. The most salient derived conclusions are, first, that Thr 109 (AATase) or Ser 109 (TATase) is an important discriminator for dicarboxylic acid selectivity by these two enzymes (I: < -2.9 kcal/mol). The T109S mutation in AATase produces a nearly equal and opposite effect to S109T in TATase (C: < 0.4 kcal/mol). Second, an I: value of 5.5 kcal/mol describes the effects of mirror mutations D94S (site 1) and S55D (site 2) in the Ca(2+) binding sites of oncomodulin on Ca(2+) affinity. The second mirror set, G98D (site 1) and D59G (site 2), yields a smaller impact (I: = -3.4 kcal/mol) on Ca(2+) binding; however, the effect is significantly more nearly context independent (C: = -0.6 versus C: = -2.7 kcal/mol). Third, the stem and loop regions of HIV and BIV TAR are predominantly responsible for the species specific interaction with BIV Tat(65-81) (I: = -1.5 to -1.6 kcal/mol), whereas I: = 0.1 kcal/mol for bulge TAR chimeras. The C: values are from -0.3 to -1.2 kcal/mol. The analysis described should have important applications to protein design.  相似文献   

16.
Kynurenine products in tryptophan metabolism are of crucial importance in modulation of neurodegenerative processes in the CNS. Kynurenic acid (KYNA): the endogenous antagonist of ionotropic glutamate receptors, displays specific affinity towards glycine site ofNMDA-receptor NR1 subunit. Mechanisms for the selective interaction of KYNA and its derivatives with other glutamate receptor subtypes are studied insufficiently. Ab initio quantum chemical calculations for KYNA-imidazole dimer, as a model for ligand interaction with His88 fragment of NR2A-subunit, along with KYNA-phenol dimer, as a model for ligand interaction with Tyr61 fragment of GluR2-subunit, were carried out in order to investigate stacking-interaction role of KYNA binding by NR2A subunit of NMDA-receptor and GluR2 subunit of AMPA-receptor. Stacking-interaction energy of KYNA-H88 for the assumed ligand orientation in the binding site is 3.0-5.0 kcal/mol and 102. kcal/mol for the optimized dimer KYNA-imidazole geometry. Stacking-interaction energy of KYNA-Tyr61 for the assumed ligand orientation in the binding site is 6.7-8.5 kcal/mol. The obtained values are comparable with the energies of hydrogen bonds. Thus, stacking-interaction should be taken into account while studing ligand glutamate receptor binding mechanisms. Stacking-interaction is evidently important for the initial ligand orientation inside the receptor binding site after which the delicate tuning of hydrogen bonding pattern is realized. On the other hand, the specific affinity of KYNA derivatives to the receptor subunits could be explained by ligand-aromatic receptor aminoacid stacking-interaction geometry difference.  相似文献   

17.
Binding of Ca(2+) to the regulatory domain of troponin C (TnC) in cardiac muscle initiates a series of protein conformational changes and modified protein-protein interactions that initiate contraction. Cardiac TnC contains two Ca(2+) binding sites, with one site being naturally defunct. Previously, binding of Ca(2+) to the functional site in the regulatory domain of TnC was shown to lead to a decrease in conformational entropy (TDeltaS) of 2 and 0.5 kcal mol(-1) for the functional and nonfunctional sites, respectively, using (15)N nuclear magnetic resonance (NMR) relaxation studies [Spyracopoulos, L., et al. (1998) Biochemistry 37, 18032-18044]. In this study, backbone dynamics of the Ca(2+)-free regulatory domain are investigated by backbone amide (15)N relaxation measurements at eight temperatures from 5 to 45 degrees C. Analysis of the relaxation measurements yields an order parameter (S(2)) indicating the degree of spatial restriction for a backbone amide H-N vector. The temperature dependence of S(2) allows estimation of the contribution to protein heat capacity from pico- to nanosecond time scale conformational fluctuations on a per residue basis. The average heat capacity contribution (C(p,j)) from backbone conformational fluctuations for regions of secondary structure for the regulatory domain of cardiac apo-TnC is 6 cal mol(-1) K(-1). The average heat capacity for Ca(2+) binding site 1 is larger than that for site 2 by 1.3 +/- 0.8 cal mol(-1) K(-1), and likely represents a mechanism where differences in affinity between Ca(2+) binding sites for EF hand proteins can be modulated.  相似文献   

18.
Prostaglandin endoperoxide H synthases 1 and 2, also known as cyclooxygenases (COXs) 1 and 2, convert arachidonic acid (AA) to prostaglandin endoperoxide H(2). Prostaglandin endoperoxide H synthases are targets of nonspecific nonsteroidal anti-inflammatory drugs and COX-2-specific inhibitors called coxibs. PGHS-2 is a sequence homodimer. Each monomer has a peroxidase and a COX active site. We find that human PGHS-2 functions as a conformational heterodimer having a catalytic monomer (E(cat)) and an allosteric monomer (E(allo)). Heme binds tightly only to the peroxidase site of E(cat), whereas substrates, as well as certain inhibitors (e.g. celecoxib), bind the COX site of E(cat). E(cat) is regulated by E(allo) in a manner dependent on what ligand is bound to E(allo). Substrate and nonsubstrate fatty acids (FAs) and some COX inhibitors (e.g. naproxen) preferentially bind to the COX site of E(allo). AA can bind to E(cat) and E(allo), but the affinity of AA for E(allo) is 25 times that for E(cat). Palmitic acid, an efficacious stimulator of human PGHS-2, binds only E(allo) in palmitic acid/murine PGHS-2 co-crystals. Nonsubstrate FAs can potentiate or attenuate actions of COX inhibitors depending on the FA and whether the inhibitor binds E(cat) or E(allo). Our studies suggest that the concentration and composition of the free FA pool in the environment in which PGHS-2 functions in cells, the FA tone, is a key factor regulating PGHS-2 activity and its responses to COX inhibitors. We suggest that differences in FA tone occurring with different diets will likely affect both base-line prostanoid synthesis and responses to COX inhibitors.  相似文献   

19.
Phosphopyridoxyl derivatives, which are stable analogues of a substrate-coenzyme complex, are bound at the active site with great affinity. From a comparison of the interaction of a number of such compounds with the apoenzyme the delta G0 values for the binding of the substrate carboxy and phenyl groups and of the coenzyme aldehydic group were determined to be equal to (or more negative than) -3.8. -8.4 and -12.5kJ/mol (-0.9, -1.9 and -3kcal/mol) respectively; the delta G0 for the binding of the coenzyme phosphate group was shown to be more negative than -20.5kJ/mol (-4.9kcal/mol). Two features of the binding process of the coenzyme-substrate analogues to tyrosine decarboxylase have already been found in the case of tyrosine aminotransferase [Borri-Voltattorni, Orlacchio, Giartosio, Conti & Turano (1975) Eur. J. Biochem. 53, 151-160]: (1) in the binding of the substrate to the enzyme a significant fraction of the instrinsic delta G0 appears to be used for some associated endoergonic process; (2) the delta H0 and delta S0 of binding appear to be very sensitive indicators of the correct alignment of the substrate-coenzyme and analogues at the active site.  相似文献   

20.
The mechanism by which the inflammatory enzyme prostaglandin H(2) synthase-1 (PGHS-1) deactivates remains undefined. This study aimed to determine the stabilizing parameters of PGHS-1 and identify factors leading to deactivation by nitric oxide species (NO(x)). Purified PGHS-1 was stabilized when solubilized in beta-octylglucoside (rather than Tween-20 or CHAPS) and when reconstituted with hemin chloride (rather than hematin). Peroxynitrite (ONOO(-)) activated the peroxidase site of PGHS-1 independently of the cyclooxygenase site. After ONOO(-) exposure, holoPGHS-1 could not metabolize arachidonic acid and was structurally compromised, whereas apoPGHS-1 retained full activity once reconstituted with heme. After incubation of holoPGHS-1 with ONOO(-), heme absorbance was diminished but to a lesser extent than the loss in enzymatic function, suggesting the contribution of more than one process to enzyme inactivation. Hydroperoxide scavengers improved enzyme activity, whereas hydroxyl radical scavengers provided no protection from the effects of ONOO(-). Mass spectral analyses revealed that tyrosine 385 (Tyr 385) is a target for nitration by ONOO(-) only when heme is present. Multimer formation was also observed and required heme but could be attenuated by arachidonic acid substrate. We conclude that the heme plays a role in catalyzing Tyr 385 nitration by ONOO(-) and the demise of PGHS-1.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号